You are on page 1of 7

Environ. Sci. Techno/.

1995, 29, 439-445

Reductive Dechlorination of Introduction


Halogenated hydrocarbons are common soil, sediment, and
Carlron Tetrachloride in Water groundwater pollutants (1-5). Many halogenated hydro-
carbons are toxic, mutagenic materials that are resistant to
Catalyzed by Mineral-Supported microbiological degradation and hence have long envi-
ronmental half-lives. The aliphatic carbon atoms bonded
Biomimetic Cobalt Macrocycles to more than two halogens or the carbon atoms in aromatic
rings with high halogen substitution have high formal
LJERKA UKRAINCZYK,*'+ oxidation states, which render such molecules resistant to
M A L A M A CHIBWE,' the degradation under aerobic conditions and susceptible
THOMAS J . PINNAVAIA,' A N D to a reductive degradation (1). This suggests the possibility
STEPHEN A . B O Y D t of using the reductive dehalogenation to detoxify haloge-
Department of Crop and Soil Sciences, and Department of nated organic compounds. Certain anaerobic micro-
Chemistry and Center for Fundamental Materials Research, organisms reductively degrade such compounds by co-
Michigan State University, East Lansing, Michigan 48824 metabolism; however, these processes are often slow and
require proper environmental conditions for microbial
activity and growth (6-8). Furthermore, high contaminant
concentrations may be toxic to the pollutant-degrading
Reductive dehalogenation, mediated by nonspecific bacteria (6, 7). In principle, the problem of toxicity can be
biomimetic Co macrocycles, was studied in aqueous overcome by using biomimetic catalysts rather than live
systems using carbon tetrachloride as a model cells. Reductive dehalogenation has been shown to be
compound. Two water-soluble macrocycles, cobalt mediated by transition metal macrocycles, such as co-
t e t r a k i s( N- m e t h y 1-4- p y r i d in iu my1) p or p h y r i n enzyme F430, and cobalamin (9-16). These macrocycles
(CoTMPyP) cation and cobalt tetrasulfophthalocyanine alongwith several other biomimetic complexes (12,13,17-
(CoPcTs) anion, were used as homogeneous and 22) have been studied in homogeneous abiotic aqueous
systems as potential remediation catalysts for biomimetic
mineral-supported catalysts. The supported catalysts reductive dehalogenation of halogenated hydrocarbons.
were prepared by exchanging CoTMPyP on the Reductive dehalogenation catalyzed by biomimetic
hectorite, fluorohectorite, and amorphous silica surface macrocycles exhibits broad substrate specificity (10-13,
and by exchanging CoPcTs on the layered double 23). Among various metallated macrocycles, those con-
hydroxide surface. Supported macrocycles were taining Co and Ni were found to be the most effective
catalytically active in the dechlorination of CC14 and catalysts (11-13). The catalytic activity of Co can be
the initial reaction rates followed Michaelis-Menten attributed to its low spin electronic configuration in various
macrocycles and to the unique property of cobalt to form
kinetics. The value of vmaxwas correlated t o the
metal-carbon bonds in water (24). The superior catalytic
previously reported orientation of macrocycles in the activity of Ni appears to be limited to coenzyme F430, which
interlayers and to the accessibility and electronic was shown to be more active than Co macrocycles (11,131,
state of the metal center, following the order: CoTMPyP- while other Ni porphyrins exhibited lower catalytic activity
silica > CoPcTs-layered double hydroxide > CoT- than Co porphyrins (12).
MPyP-fluorohectorite > CoTMPyP-hectorite. In both The implementation of homogeneous macrocyclic cata-
heterogeneous and homogeneous systems, volatile lysts in remediation technologies is impractical because of
reaction products accounted for less than 30% of CC14 problems with separating the aqueous catalyst, reactants,
and products. This problem could be overcome by
degraded. In short-term experiments (2 h), homo- immobilizing macrocyclic complexes on a solid support.
geneous CoTMPyP was more active than heterogeneous There are numerous examples of supported macrocycles
catalysts, while homogeneous CoPcTs was deacti- and their applications in heterogeneous catalysis;however,
vated due to aggregation, and degraded less CC14 than their application in aqueous systems has rarely been
supported CoPcTs. In long-term experiments (3 investigated (22). Recently photoreductive dehalogenation
days), where large CC14/macrocycle ratios were used, of aqueous bromoform has been shown to be catalyzed by
silica-supported CoTMPyP was more active than the anionic macrocycle (CoPcTs)adsorbed on the positively
charged titania surface (21).
homogeneous CoTMPyP, suggesting that adsorption ,
High surface area minerals, such as layered clay minerals
stabilized the catalyst. and colloidal oxides, are attractive potential catalyst sup-
ports for environmental remediation in aqueous systems
from both an economical and practical point of view.

* Address correspondence to this author at his present address:


Department of Agronomy, Iowa State University, Arnes, IA 50011;
e-mail address: I-ukrain @ iastate.edu.
t Department of Crop and Soil Sciences.
*Department of Chemistry and Center for Fundamental Materials
Research.

0013-936W95/0929-0439$09.00/0 @ 1995 American Chemical Society VOL. 29. NO. 2. 1995 / ENVIRONMENTAL SCIENCE & TECHNOLOGY 439
H3C- -CH3

CH3

COTMPyP CoPcTs
FIGURE 1. Structure of the cobalt tetrakis(l-metfiyl-4-pyridiniumyl)po~hyrin(CoTMPyP) cation and cobalt tetrasulfophthalocyanine(CoPcTs)
anion.

Investigation of these systems is of interest for remediation CoTMPyP-hectorite a


of industrial effluents and groundwater and for in situ
remediation of soils and sediments. The macrocycle must
carry a permanent charge to be adsorbed and retained on
II f
the charged surfaces of these supports. Synthetic cationic
p
co- -co- I 14.OA
and anionic porphyrins and phthalocyanines have been
intercalated into aluminosilicate clays and layered double
P
hydroxides (25-35). Unlike most neutral porphyrins (36-
381, cationic porphyrins are stable against demetallation CoTMPyP-fluorohectorite b
on the acidic clay surfaces (28).
Two macrocyclic catalysts, cationic CoTMPyP and
anionic CoPcTs (Figure l),were used in this study. These J
I 1
macrocycles have been previously intercalated in layered
minerals and characterized (35)byX-ray diffraction (XRD),
/////
o c o o c o OCO 0 c o 0 c o o
19.6A
electron spin resonance (ESR), and ultravioletlvisible ( U V / / / / / A 270
vis) spectroscopy. Their arrangements in the galleries of
layered supports are presented in Figure 2.
m
Carbon tetrachloride (CT) was chosen as a model
compound to study dehalogenation reactions by the C
supported macrocyclic catalysts because its biomimetic CoPcTs-layered double hydroxide
dechlorination in homogeneous systems has been studied I I
extensively. Although there are a number of studies on
reductive dechlorination of CT in water catalyzed by cobalt
macrocycles (10-1 4, la,the mechanismsof these reactions
are still not fuuy understood (16). Reaction schemes
1 l l l I I l
o c o o c o o c o o c o o coo c o o c o o 23.7A
involving Co(I1) (16) and Co(1) (14) as a catalytically active
species and the alkyl-cobalt intermediates have been
proposed. The volatile reaction products have been
identified, usually accounting from 20-30% (10,16)up to
50% (10,141of CT degraded. The major volatile degradation FIGURE 2 Arrangement of CoTMPyP in the interlayers of (a) hectorite
products are chloroform (CF) and dichloromethane (DM), and (b)fluorohectorite and of (c) CoPcTs in the interlayersof layered
double hydroxide (according to ref 35).
while usually only trace amounts of chloromethane,
methane, and carbon monoxide have been detected (10, The objectives of this study were to investigate whether
11, 14, 16). The homogeneous dechlorination reactions cationic and anionic Co macrocycles supported on high
have usually been performed in buffered solutions using surface area minerals maintain their ability to catalyze
dithiothreitol (DTT; 12,15, l a , titanium(II1) citrate (10,11, reductive dechlorination, to compare the rates of biomi-
13, 141, cysteine (191, sodium dithionate (201, or sodium metic reductive dehalogenation by homogeneous and
disulfide (19) as a reductant. Sodium dithionate was not heterogeneous macrocycles, and to investigate the long-
used in this study because it produces free radicals that term stability of supported catalysts.
destroy CoTMPyP (39). Sodium disulfide was used in the
preliminary experiments; however, difficulties were en- Materials and Methods
countered in buffering the pH. The experiments reported AU chemicals were reagent grade unless otherwise indicated
here were conducted using DTT as a reductant. and were used without further purification. CoTMPyP was

440 1 ENVIRONMENTAL SCIENCE & TECHNOLOGY / VOL. 29, NO. 2,1995


obtained as the tetrachloride salt from Midcentury Co, was less than 1%in all short-term experiments (after 2 h)
(Posen, IL). CoPcTs was synthesized as the tetrasodium and less than 4% in all long-term experiments (after3 days).
salt by a published procedure (30,40). Natural hectorite To test whether the dechlorination is catalyzed by the
[M(Uo,dMg5,31Li0,1d(S$Ozo(OH,F)zl from Hector, CA, was complex at the interface, the reacting suspension in short-
obtained in a spray-dried form from the Baroid Division of term experiments was filtered after approximately 30 min
NL Industries, The CEC of this clay was 0,7 mmok g-1(41), in the glovebox through a 0.1-pm Teflon filter into another
Synthetic fluorohectorite [Li1.&4&1.&1.6) (Sh)O2oF41 was vial containing 0.015 g of solid DTT. The vials were sealed,
prepared by crystallization of a fluorohectorite glass of and the degradation of CT was followed for 2 h. Although
equivalent composition according to the general procedure the filtrate was weakly colored (indicating some of the
by B e d (42). The CEC of this material was 1,9 mmolc g-l macrocycle was desorbed from the support and/or some
(43). Synthetic hydrotalcite-like LDH of the type [Mgdz- of the colloidal particles passed through the filter),less then
(OH)14][CO3]x yHz0 was prepared bythe method ofReichle 2% of CT was dechlorinated in a 2-h time period, implying
etal. (44). Silica gel (chromatographic grade, Aldrich)with that the macrocycle did not leach into liquid phase in a
a surface area of 500 m2 g-' was used without further significant amount.
modifications. The kinetics of CT degradation and the appearance of
Preparation of Supported Catalysts. Preparation of CF and DM was followed by headspace analysis using a
layered catalystshas been described previously (35). Silica- Varian gas chromatograph with a DB-625 (0.25 mm i.d. x
supported CoTMPyP was prepared by reacting 1 g of silica 30 m, J&W Scientific) column and a flame-ionization
gel suspended in 50 mL of water with 1 x mmol of detector. The column temperature was 35 "C, injector
CoTMPyP for 24 h at pH 7.5 (adjusted with NaOH) under temperature was 200 "C, and detector temperature was
nitrogen. The suspension was then centrifuged giving 250 "C. Headspace was sampled manually using a 0.1-mL
colorless supernatant, indicating that all CoTMPyPhas been gas-tight syringe. The peak identification and quantifica-
adsorbed. The product was washed three times with water tion were done using external standards made in 0.1 M
and freeze-dried. The Soret band of suspended CoTMPyP- phosphate buffer. In the preliminary experiments, it was
SG was at 438 nm, and the complex was ESR silent, determined that the peak areas of the standards were not
indicating that Co was oxidized to Co(II1). affected by the presence of the solid catalyst in the amount
DechlorinationExperiments. Dechlorinationreactions used in this study.
were studied using a modified method of Krone etal. ( 1I).
The reaction was carried out in a 2-mL anaerobic mixture Results and Discussion
placed in a 10-mL glass vial crimp-sealed with a Teflon- Short-TermExperiments. Kinetic data for CT degradation
coated rubber septum (Hewlett-Packard)and wrapped with in six representative systems (homogeneous CoTMPyP,
aluminum foil. The vials were filled and sealed in an CoTMPyP-FH, CoTMPyP-H,CoTMPyP-SG, homogeneous
anaerobic glovebox (Plas-Labs) under 10% H2/90% N2 CoPcTs, CoPcTs-LDH) are shown in Figure 3. All of the
atmosphere. In short-term experiments (reaction time is supported macrocycles were catalytically active. CF and
2 h), the catalyst/reductant molar ratio was 1:lOOO (0.1mM DM were detected by the headspace analysis as the
CoTMPyP or CoPcTslO.1 M DTT), and in the long-term degradation products in homogeneous systems,while only
experiments (3 days) this ratio was 1:25000 (0.002 mM CF was detected in the heterogeneos systems even at
CoTMPyP or CoPcTsl0.05 M DTT). reaction times greater than 2 h. The volatile products
In short-term experiments, the homogeneous catalyst accounted for less than 30% of CT degraded, which is
was prepared as a 0.01 M stock solution and added by a consistent with previous studies of CT degradation medi-
microsyringe. The heterogeneous catalyst was added by a ated by c o macrocycles (10, 1 1 , 13, 16, 19).
100-pL pipet as an aqueous suspension (0.55 wt % suspen- The heterogeneous degradation of CT catalyzed by Co
sion of CoTMPyP-FH, 1.03wt % suspension of CoTMPyP- macrocycles illustrated in Figure 3 can be described by the
H, 0.34 wt % suspension of CoPcTs-LDH). CoTMPyP-SG first-order kinetics:
was added as a solid (0.020@. Dithiothreitiol(Mallinckrodt)
was added as a solid. Freshly prepared, deoxygenerated,
and autoclaved 0.1 M sodium phosphate buffer (pH 7.5)
was added by a pipet, and the vials were sealed. Reaction
was initiated by injecting freshly prepared CT (99.9%,
The pseudo-first-order rate constants (kobs)were calculated
Burdick & Jackson) stock solution (0.062MI in 2-propanol
(Table 1) by fitting the integrated form of the eq 1 to the
through a septum 5 min after assembling the reductant/
data in Figure 3. The observed pseudo-first-order rate
catalystlbuffer mixture. The vials were shaken on a
constant (kobs) may conceptually be separated into two
controlled temperature rotary shaker at 25 "C for a desired
constants: kl, a pseudo-first-order rate constant for the
period of time.
formation of CF, and k2, a pseudo-first-order rate constant
In long-term experiments, the catalyst was added by a
for the formation of unidentified products (includinglong-
microsyringe (homogeneous) or pipet (heterogeneous) to
75 mL of the phosphate buffer, and 2 mL of this mixture
lived intermediates): kobs = kl +
k2. Then the rate of
appearance of CF is
was pipetted into 10-mL glass vials. The other reactants
were added as in short-term experiments. The experiments
were performed in duplicate, and each vial was analyzed
only once. One control, containing CT and DTT in
phosphate buffer, was run for each time point. Two random
controls were runas well: one containing CT and supported
catalyst in phosphate buffer, and the other containing CT where k3 is a pseudo-first-order rate constant for CF
and the phosphate buffer. The loss of CT from the controls degradation. The rate of formation of DM is

VOL. 29, NO. 2. 1995 / ENVIRONMENTAL SCIENCE & TECHNOLOGY I 4 4 1


aggregation (30, 331, and thus the rate of heterogeneous
degradation does not show the rapid decrease in activity.
At the same total catalystlCT molar ratio (l:lO), the
0.6 1\ 1 degradation rates in CoTMPyP systemswere slower for the
0.4 heterogeneous systems than the homogeneous system,
0.2

0.0
0 30 60 90 120
O.* 0
0.0 u 30 60 90 120
following the order: CoTMF'yP CoTMPyP-SG CoT-
MPyP-FH > CoTMPyP-H. Adsorption on clay surfaces has
often been shown to alter the reactivity, mobility, and redox
properties of transition metal complexes and enzymes (41,
1.o 1 .o 43, 46, 4 7)
I

C.
0.8
Reactions catalyzed by macrocycles at the clay or silica
surface can differ from homogeneous reactions as a result
0.6
of changes in the redox potential of the supported mac-
0.4 rocycles as well as a result of the geometricalarrangements
0.2 0.2
ofthe macrocycles at the interface, which may yield different
product distribution. The lack of formation of DM in the
0"E 0.0
0 30 60 90 120
0.0
0 30 60 90 120 heterogeneous systems in this study is most likelythe result
of a change in the reduction potential of the supported
macrocycles to a more positive value relative to the aqueous
macrocycles. Since the dechlorination of CF to DM has
lower reduction potential [B&FIDM) = f0.439 V at pH
7, calculated from free energies of formation in liquid state,

0.2
0'4
0.0 w
0 30 60 90 120
I 0 30 60 90 (20
1 M C1- aqueous solution; 11,481 than the dechlorination
ofCT to CF [Bo(CTICF)= +0.516Vl, thereduction potential
of a supported Co macrocycle may be too high to drive the
reduction of CF to DM.
Furthermore, in solution both axial positions on Co are
Time (min) available for reaction while in the supported system the
FIGURE 3. Degradation of 1.0 mM CT (0)and production of CF (0) reactants may only have access to one of the Co axial
and DM (v) at pH 7.5 catalyzed by 0.1 mM Co macrocycle: (a) positions. The quantitative information on the changes in
homogeneous CoTMPyP, (b) CoTMPyP-FH, (c) CoTMPyP-H, (d) product distribution can be obtained from the ratio of the
ColMPyP-SG, (e) homogeneous CoPcTs, and fl) CoPcTs-U)H. Cutves rate constant (Table 1) for the formation of CF (kl) to the
represent simulation of the CT dechlorination and CF and DM
formation, based on the kinetic model presented in eqs 1-3. overall rate constant for CT degradation (kobs): the smaller
this ratio, the lower the tendency for the formal two-electron
TABLE 1 transfer from CT to Co and for CF formation. Alkyl radicals
('CC13 and 'CHCl3) or carbanions formed from CT and CF
Pseude=First=OrderRate Constants ( x 103)for by one electron reduction (10)instead of donating another
Degradation of CT and Product Formation Catalyzed electron to Co and reacting with the hydrogen ion to form
by Hoatogeneous and Heterogeneous Macrocycles CF and DM may be reacting with water to give CO and
hbs ki k2 4 k, formic acid as well as reactingwith each other and with the
catalyst (min-l)O SEb (inin-')' (min-')' (min-l)' ( m i d ) ' organic reductant to give nonvolatile products (10,11, 14,
CoTMPyP 52.2 4.8 26.5 25.7 7.9 5.8
16).
CoTMPyP-FH 8.0 0.2 1.5 6.5 1.1 ndc Plot of the initial rate (& = ko[CC&lo)of CT degradation
CoTMPyP-AS 31.4 2.3 9.9 21.5 1.2 nd by the supported macrocycles as a function of the initial
CoTMPyP-H 6.4 1.0 0.6 5.8 1.0 nd concentration of CT ([CC&l0)is shown in Figure 4. The
CoPcTs 45.8d 7.5 nfe nf nf nd initial rate constants (b) for heterogeneousCT degradation,
CoPcTs-LDH 28.9 1.3 12.8 16.1 4.3 nd
calculated by fitting the first-order rate equation to the data
a Rate constants kbs, k,,k2,and k, obtained by fitting eqs 1-3 to the for the first 30 min of the experiment, decrease in the order
data in Figure 3. Standard error for k b r . CH2C12was not detected.
Initial rate constant. Not fitted because the reaction could not be COTMQP-SG> COPCTS-LDH > COTMQP-FH> COTMQP-
described by the first-order kinetics. H.
It can be assumed that in the initial stages of the reaction,
the concentration of CT is much higher than the concen-
(3) tration of any of the reaction products, and thus the product
competition for the reactive sites can be neglected. At-
where k4 is a pseudo-first-order rate constant for the tempts to measure the amount of CT adsorbed in the
formation of DM. Equations 1-3 were solved by the suspensions containingonly the supported catalyst without
Runge-Kutta method using the k&s from the Table 1 and the reductant were unsuccessful because the amount
fitting k2, k3 and k4 values to the data (Table 1). The adsorbed was too small. Apparently the decrease in CT
calculated curves are shown in Figure 3 for all the systems concentration in the heterogeneous systems can be at-
studied, except homogeneous CoPcTs. In the homoge- tributed solely to its chemical degradation. The reductant
neous CoPcTs system,the fast initial rate rapidly decreases is then necessary in order to generate catalytically active
(Figure3e),indicatingthat the catalyst is beinginactivated. Co species.
The fast decrease in the catalytic activity of this macrocycle Spectroscopicstudies have shown that Co is present as
has been attributed to the aggregation of CoPcTs in aqueous Co(I1) in the intercalated macrocycles under both 0 2 and
solutions (45). Intercalation stabilizes CoPcTs against NZatmosphere (35). Since the supported catalyst exhibits

442 ENVIRONMENTAL SCIENCE & TECHNOLOGY / VOL. 29, NO. 2,1995


61 I I I I I 1

1 i
coTMP
COPcTs-LDH

R (mMh')

0 0.5 1 1.5 2 2.5 3

[CCI~IO
(mM)
FIGURE 4. Plot of the initial rate of degradation of CT in heterogeneous systems as a function of the initial CT concentration (pH 7.5,O.l
mM catalyst, 0.1 M DTr). Curves represent the Michaelis-Menten model (eq 4).

TABLE 2 order rate equation. The value of ,,v (Table 2) is


proportional to the surface density of catalytically active
Michaelis-Menten Parameters for CT Degradation sites, following the order CoTMPyP-SG > CoPcTs-LDH >
by Heterogeneous Catalysts CoTMPyP-FH > CoTMPyP-H. For layered supports, this
catalyst rU0 (mMl v, (mM h-? order is in good agreementwith the results from the previous
spectroscopic studies on the orientation and hydration of
CoTMPyP-FH 4.48 6.60
CoTMPyP-SG 5.87 17.28
the macrocycles in the interlayers (Figure 2; 35).
CoTMPyP-H 2.34 1.OB The lowest reaction rates were observed for the CoTMPyP
COPCTS-LDH 5.41 10.41 supported on the low charge density clay hectorite, where
the macrocycle is lying parallel to the clay layers. The ESR
no activity in the absence of the reductant, the reduction spectra of Co(I1)TMPyP intercalated in hectorite clearly
of Co(I1) to Co(1) by DTT may have to take place in order show that Co has no water in the axial coordination sites
to produce a catalytically active Co species. Adsorption of and its d orbitals, interactingwith electron-deficient siloxane
CT onto the catalyst and the formation of the Co-CC13 oxygens, are contracted (33, indicating that it is likely to
intermediate in the heterogeneous systems depends on be kinetically inert in electron-transferreactions (51). Water
the number of reduced Co sites and their accessibility to does not diffuse into the CoTMPyP-H interlayers when the
the reactants. The decrease in k~ with increased CT clay is fullywetted, and apparently reactants cannot access
concentration (Figure 4) indicates that there is a limited the interlayers either. Only a small portion of the Co(I1)-
number of easily accessible active sites. The concentration TMPyP, adsorbed on the clay edges and outer surfaces, has
of the intermediate complex at the initial stages of the water in axial coordination sites, and those are most likely
reaction is small compared to the concentrations of Co the only catalytically active sites in CT dechlorination.
macrocycle and CT. Thus, the dechlorination reaction in Intercalated CoTMPyP in a high-charge density clay fluo-
heterogeneous systems is very similar to enzyme-catalyzed rohectorite is lying at a 27" angle to the clay layers with
reactions, and the initial rate of CT degradation (Figure 4) water in the axial coordination sites. The electronic
can be fitted by the Michaelis-Menten kinetic model: configuration of aqueous CoTMPyP and CoTMPyP-FH are
very similar ( 3 3 ,and the ESR study shows that water diffuses
easily in and out of the clay interlayers. Therefore, the
(4) decrease in activity for CoTMPyP-FH compared to the
aqueous CoTMPyP can be attributed to the slower rates of
where ,v is the maximum reaction rate for a specified adsorption of CT on active metal centers in the interlayers.
initial Co macrocycle (enzyme) concentration and Km is This may be due to the restricted mobility of reactants in
Michaelis constant. the clay interlayers as well as to the smaller number of
The value of 'K (Table 2) greater than 1implies that the reduced metal centers because of the restricted access of
rate of formation of the intermediate complex is slower DTT to the macrocycles in the interlayers. Silica-supported
than the rate of desorption of the products in all the CoTMPyP exhibits the highest activity among supported
supported systems (49, 50). At longer reaction times, CT macrocycles. Metal centers of the macrocycles supported
concentration decreases, while CF and other reaction on the spherical silica particles may be more active than
products compete for the active surface sites, which on the other supports because they are more accessible to
increases K,. When K, >->[CCL],eq 4 reduces to the first- the reductant and to the reactants and possibly because

VOL. 29, NO. 2, 1995 I ENVIRONMENTAL SCIENCE & TECHNOLOGY 448


TABLE 3
Pserdo=First=OrderRate Constants for CT
Degradation and Turnover Frequencies in Long=Term
Dechlorination Experiments
::b
0.0
-
j 0.8

catalyst
ICCl410
(mM)
k
(day-')* SEb
turnover
frequency (day-')c
0.0 0.5 1.0 1.5 2.0 2.5 3.0 0.0 0.5 1.0 1.5 2.0 2.5 3.0
CoTMPyP 2.3 0.49 0.06 570
CoTMPyP-FH 2.3 0.47 0.08 540
COTMPyP-SG 2.3 0.57 0.07 655
CoTMPyP-H 2.3 0.09 0.01 105

-e
16 CoPcTs 2.3 0.75 0.09 860
COPCTS-LDH 2.3 0.29 0.67 335
CoTMPyP 4.6 0.28 0.06 640
0.8 CoTMPyP-SG 4.6 0.39 0.04 895
8 0.4
0.0
0.0 0.5 1.0 1.5 2.0 2.5 3.0
0.0
0.4 lLE!%zd
0.0 0.5 1.0 1.5 2.0 2.5 3.0
*
*Calculated forthefirst24 h ofexperioments. Standarderror. 0 Mol
of CCl,mol of catalyst-' day-'calculated for the first 24 h of experiment.

1.6
12

0.0 0.5 1.0 1.5 2.0 2.5 3.0


!im
12
0.8

0.4
0.0

Time (days)
i_,vr""i
0.0 0.5 .1.0 1.5 2.02.5 3.0
the slower rates can be attributed to the slow rates of
adsorption of CT on reactive porphyrin metal centers in
the interlayers rather than to the loss of catalyst longevity
of the macrocycle upon intercalation. Low turnover
frequency exhibited by hectorite-supported CoTMPyP may
be attributed in the inaccessibility of the intercalated
porphyrin and to the alteration of the electronic state of
the metal center (35).
The hydration and electronic state of Co(1I)PcTs inter-
calated in LDH are similar to that of CoTMPyP-FH. This
FIGURE 5. Degradation of 23 mM CT (0)and production of CF (0)
at pH 7.5 catalyzed by 0.002 mM 60 macrocycle: (a) homogeneous macrocycle is perpendicular to the layers of LDH, providing
CoTMPyP, (b) CoTMPyP-FH, (c) CoTMPyP-H, (d) CoTMPyP-SG, (e) a more open structure and thus more easily accessible metal
homogeneous CoPcTs, and (f) CoPcTs-LDH. center as compared to CoTMPyP-FH, which may partly
explain the higher dechlorination rates. In addition, the
the reduction potential of Co is less altered on this surface difference in dechlorination rates between CoTMPyP-FH
as compared to highly charged layered mineral surfaces. and CoPcTs-LDH may partly arise from the difference in
Long-Term Experiments. The effect of immobilization the catalytic activity of two macrocycles in CT dechlori-
of the macrocycles on the solid support on the loss of their nation exhibited in the homogeneous systems (Figure 5a
catalytic activity with time was investigated in long-term and d; Table 3).
experiments (Figure5). The experiments were performed
with low catalyst amounts (0.002 mM) and high CT Corclusions
concentrations (2.3 and 4.6 mM). The homogeneous
CoPcTs catalystwas not deactivatedunder those conditions, The study demonstrates that the supported macrocycles
because the aggregation of CoPcTs does not occur in very are catalytically active under ambient conditions. The
dilute solutions. After 1 day, the reaction did not follow heterogeneous dechlorination reactions can be described
the first-order kinetics because the catalysts had lost some by the first-order kinetics with the pseudo-first-order rate
of their activity. Thus, the first-order rate constants and constants being correlated to the accessibility and the
turnover frequencies (mole of CT degraded per mole of hydration of the cobalt in supported macrocycles, following
macrocycle per day) were calculated for the reaction time the order: CoTMPyP-SG > CoPcTs-LDH > CoTMPyP-FH
of 1 day (Table 3). > CoTMPyP-H. The reaction mechanism may be similar
Among the supported catalysts, CoTMPyP-SG exhibited to enzyme-catalyzedreactions where the rate-determining
the highest activity, and CoTMPyP-H exhibited the lowest step is adsorption of CT on the active center with the
activity. The frequency turnover was greater for CoTMPyP- formation of the intermediate complex. Although initial
SG than for the homogeneous CoTMPyP (Table 31, indicat- heterogeneous dechlorination rates were less than the
ing that adsorption on silica surface has significantly corresponding homogeneous rates, increased longevity of
increased catalyst longevity. For other supports (H, FH, supported catalysts led to greater dechlorination in reac-
and LDH), turnover frequencies were less than those of the tions catalyzed by CoTMPyP-SG and CoTMPyP-FH than
corresponding homogeneous catalyst. With the exception by homogeneous CoTMPyP at longer reaction times.
of hectorite, the loss of catalytic activity of the Co mac- The charged, mineral-supported macrocycles reduc-
rocycles does not seem to be promoted by adsorption on tively dechlorinate carbon tetrachloride in water, even at
the solid. In fact, the homogenous CoTMPyP has lost its high concentrations that would inhibit microbial activity.
activity after 2.5 days while the CoTMPyP-FH and CoT- Since the reduction potential of the macrocycles can be
MFyP-SG systems were still active. altered by choosing the appropriate substituents on the
The smaller turnover frequency of CoTMPyP-FH than periphery of the porphyrin or phthalocyanine ligand,
that of homogeneous catalyst over the first 24 h of efficient supported catalysts could be designed for reductive
experiment is consistent with the decreased rates observed dehalogenation by synthesizing macrocycles with tailored
in short-term experiments for supported macrocycle. Thus, redox properties.

444 rn ENVIRONMENTAL SCIENCE & TECHNOLOGY / VOL. 29, NO. 2, 1995


Acknowledgments (25)Kaneyama, H.;Suzuki, H.; Amabo, A. Chem. h t t . 1988,1117.
(26) Carrado, K. A,; Winans, R. E. Chem. Muter. 1990,2,328.
This research was supported by the National Institute for (27)Giannelis, E. P. Chem. Mater. 1990,2,827.
EnvironmentalHealth Sciences Grant P42EF0491, Michigan (28) Carrado, K.A.; Thiyagarajan, P.; Winans, R. E.; Botto, R. E. Inorg.
State University Institute for Environmental Toxicology, Chem. 1991,30,794.
(29)Gaillon, L.; Bedioui, F.; Devynck, J.; Battioni, P.; Barloy, L.;
and Michigan Agricultural Experiment Station. We thank Mansuy. D. J. Ekctroanal. Chem. 1991,303,283.
Dr. Max M. Mortland for helpful discussions. (30) Perez-Bemal, M. E.; Ruano-Casero, R.; Pinnavaia, T. J. Cutal.
Lett. 1991,11, 55.
Literature Cited (31)Barloy, L.; Lallier, J. P.; Battioni, P.; Mansuy, D.; Piffard, Y.;
Tournoux, M.; Valim, J. B.; Jones, W. New J. Chem. 1992,16,71.
(1) Vogel, T. M.; Criddle, C. S.; McCarty, P. L. Environ. Sci. Technol. (32)Carrado, K. A,; Forman, J. E.; Botto, R. E.; Winans, R. E. Chem.
1987,21, 722. Mater. 1993,5, 472.
(2) Kuhn, E. P.; Suflita, J. M. In Reactions and Movement of Organic (33)Chibwe, M.;Pinnavia, T. J. J. Chem. SOC.Chem. Commun. 1993,
Chemicals in Soik;; Sawhney, B. L., Brown, K., Eds.; Soil Science 278-280.
Society of America Special Publication 22;Soil Science Society (34)Gaillon, L.; Bedioui, F.; Devynck, J.; Battioni, P. J. Elechoanal.
of America: Madison, WI, 1989;pp 111-180. Chem. 1993,347,435.
(3)Furukawa, K. In Biodegradation and detoxification of organic (35)Ukrainczyk, L.; Chibwe, M.; Pinnavaia, T. J.; Booyd, S. A. J. Phys.
pollutants;Chakrabarty, A. M., Ed.; CRC Press, Inc.: BoccaRaton, Chem. 1994,98,2668.
FL, 1982;pp 33-57. (36)VanDamme, H.; Crespin, M.; Obrecht, F.; Cruz, M. I.; Fripiat, J.
(4)Schwille, F. Dense chlorinated solvents in porous andfractured J.J. Colloid Interface Sci. 1978,66,43.
media, 1st ed.; Lewis: Chelsea, MI, 1988;146 pp. (37)Abdo, S.;Cruz, M. I.; Fripiat, J. J. Cluys ClayMiner. 1980,25,125.
(5) Scheunert, I. In Fate and prediction ofenvironmental chemicals (38)Bergaya, F.;Van Damme, H. Geochim. Cosmochim. Acta 1982,
in soils, plants and aquatic systems, 1st ed.; Mansour, M., Ed.; 46,349.
Lewis: Bocca Raton, FL, 1993;pp 1-22. (39)Pasternack, R. F.; Cobb, M. A. Sutin, N. Inorg. Chem. 1975,14,
(6) Mohn, W. W.; Tiedje, J. M. Microbiol. Rev. 1992,56, 482. 866.
(7) Morris, P. J.; Quensen, J. F., 111; Tiedje, J. M.; Boyd, S. A. Environ. (40)Weber, J. H.; Busch, D. H. Inorg. Chem. 1965,4,469.
Sci. Technol. 1993,27,1580. (41)Berkheiser, V. E.; Mortland, M. M. Cluys Clay Miner. 1977,25,
(8) Mikesell, M.D.; Boyd, S. A. Appl. Environ. Microbiol. 1990,56, 105.
1198. (42)Beall, G. H. US. Patent 3 756 838,1973.
(9)Bieniek, D.; Mosa, P. N.; Klein, W.; Korte, F. Tetrahedron Lett. (43)Newsham, M. D.; Giannelis, E. P.; Pinnavia, T. J.; Nocera, D. G.
1970,47,4055. J. Am. Chem. SOC.1988,110, 3885.
(10)Krone, U.E.; Thauer, R. K.; Hogenkamp, H. P. C. Biochemistry (44)Reichle, W. T.; Kang, S. Y.; Everhardt, D. S. J. Catul. 1986,101,
1989,28,4908. 352.
(11)Krone, U. E.; Laufer, K.; Thauer, R. K.; Hogenkamp, H. P. C. (45)Yang, Y.C.; Ward, J. R.; Seiders, R. P. Inorg. Chem. 1985,24,1765.
Biochemistry 1989,28,10061. (46)Traynor, M. F.; Mortland, M. M.; Pinnavaia, T. J. Clays ClayMiner.
(12)Marks, T. S.;Allpress, J. D.; Maule, A. Appl. Environ. Microbiol. 1978,26,318.
1989,55, 1258. (47)Boyd, S.A.; Mortland, M. M. In Soil biochemistry, 1st ed.; Bollag,
(13)Gantzer, C. J.; Wackett, L. P. Environ. Sci. Technol. 1991,25,715. J. M.; Stotzky, Eds.; Marcel Dekker: New York, 1990 Vol. 6,pp
(14)Krone, U. E.; Thauer, R. K.; Hogenkamp, H. P. C.; Steinbach, K. 1-28.
Biochemistry 1991,30, 2713. (48)Wagman, D. D.; Evans, W. H.; Parker, V. B.; Halow, L.; Bailey,
(15)Assaf-Amid,N.; Nies, L.; Vogel, T. M. Appl. Environ. Microbiol. S. M.; Schumm, R. H. NBS Technical Note; United States
1992,58, 1057. Department of Commerce, National Bureau of Standards:
(16)Assaf-Amid,N.; Hayes, K. F.; Vogel, T. M. Environ. Sci. Technol. Washington, DC, 1968;Vol. 270,p 3.
1994,28,246. (49)Spiro, M. In Reactions at the solid-liquid interface; Compton,
(17)Miskus, R.P.; Blair, D. P.; Casida, J. E. J. Agric. Food Chem. 1965, R. G., Ed.; Elsevier: New York, 1989;pp 69-107.
13, 481. (50)Campbell, I. M. Catalysis at surfaces; Chapman and Hall: New
(18)Zoro, J. A.; Humter, J. M.; Eglinton, G.; Ware, G. C. Nature York, 1988;pp 11-51.
(London) 1974,247,235. (51)Purcell, K. F.; Kotz, J. C. Inorganic chemistry, 2nd ed.; Saunders:
(19)Klecka, G. M.; Gonsior, S. J. Chemosphere 1984,3, 391. Philadelphia, PA, 1977;pp 530 and 669-680.
(20)Baxter, R. M. Chemosphere 1990,21, 451.
(21)Kuhler, R. J.; Santo, G.A.; Caudd, T. R.; Betterton, E.A.;Arnold,
R. G. Environ. Sci. Technol. 1993,27,2104. Received for review May 10, 1994. Revised manuscript re-
(22) Maldotti, A,; Amadelli, R.; Bartocci, C.; Carassiti, V.; Polo, E.; ceived November 3, 1994. Accepted November 9, 1994.@
Varani, G. Coord. Chem. Rev. 1993,125, 143.
(23)Wade, R. S.; Castro, C. E. J. Am. Chem. SOC.1973,95,226. ES9402821
(24)Cotton, F.A.; Wilkinson, G. Advanced inorganic chemistry Wiley
Interscience: New York, 1988;p 1569. @ Abstract publishedin AdvanceACSAbstructs, December 15,1994.

VOL. 29, NO. 2, 1995 / ENVIRONMENTAL SCIENCE & TECHNOLOGY 1445

You might also like